accepted manuscript - California Institute of Technology

 Specific Inhibition of p97/VCP ATPase and Kinetic Analysis Demonstrate
Interaction between D1 and D2 ATPase domains
Tsui-Fen Chou, Stacie L. Bulfer, Conrad C. Weihl, Kelin Li, Lev G.
Lis, Michael A. Walters, Frank J. Schoenen, Henry J. Lin, Raymond J.
Deshaies, Michelle R. Arkin
PII:
DOI:
Reference:
S0022-2836(14)00272-1
doi: 10.1016/j.jmb.2014.05.022
YJMBI 64466
To appear in:
Journal of Molecular Biology
Received date:
Revised date:
Accepted date:
6 February 2014
21 May 2014
22 May 2014
Please cite this article as: Chou, T.-F., Bulfer, S.L., Weihl, C.C., Li, K., Lis, L.G., Walters, M.A., Schoenen, F.J., Lin, H.J., Deshaies, R.J. & Arkin, M.R., Specific Inhibition
of p97/VCP ATPase and Kinetic Analysis Demonstrate Interaction between D1 and D2
ATPase domains, Journal of Molecular Biology (2014), doi: 10.1016/j.jmb.2014.05.022
This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
ACCEPTED MANUSCRIPT
Specific Inhibition of p97/VCP ATPase and Kinetic Analysis Demonstrate
PT
Interaction between D1 and D2 ATPase domains
Tsui-Fen Chou1* and Stacie L. Bulfer2, Conrad C. Weihl3, Kelin Li4, Lev G. Lis5, Michael
Division of Medical Genetics, Department of Pediatrics, Harbor-UCLA Medical Center and Los
NU
1
SC
Arkin2
RI
A. Walters5, Frank J. Schoenen4, Henry J. Lin1, Raymond J. Deshaies 6,7, and Michelle R.
Angeles Biomedical Research Institute, Torrance, California 90502, USA
Department of Pharmaceutical Chemistry, Small Molecule Discovery Center, University of California,
MA
2
San Francisco, California 94158, USA.
3
TE
D
Department of Neurology, Washington University School of Medicine, St. Louis, MO
4
63110, USA
5
AC
CE
P
University of Kansas Specialized Chemistry Center, University of Kansas, Lawrence, Kansas 66047,
USA
Department of Medicinal Chemistry, Institute for Therapeutics Discovery and
Development, College of Pharmacy, University of Minnesota, Minneapolis, MN 55414, USA
6
Division of Biology, California Institute of Technology, Pasadena, California 91125, USA.
7
Howard Hughes Medical Institute, USA.
*To whom correspondence should be addressed: Tsui-Fen Chou, Harbor-UCLA Medical Center and Los
Angeles Biomedical Research Institute, Torrance, California 90502, USA. Tel: 1-424-201-3006; Email:
[email protected]
Keywords: p97/VCP AAA ATPase;IBMPFD/ALS; steady-state kinetics; SPR; p97 inhibitor
Abbreviations: ERAD, endoplasmic reticulum associated degradation; SPR, surface plasmon resonance;
UFD, ubiquitin fusion degradation; UPS, ubiquitin-proteasome system; WT, wild type.
1
ACCEPTED MANUSCRIPT
ABSTRACT
The p97 AAA (ATPase associated with diverse cellular activities), also called VCP (valosin-containing
PT
protein), is an important therapeutic target for cancer and neurodegenerative diseases. p97 forms a hexamer
composed of two AAA domains (D1 and D2) that form two stacked rings, and an N-terminal domain that
RI
binds numerous cofactor proteins. The interplay between the three domains in p97 is complex, and a deeper
SC
biochemical understanding is needed in order to design selective p97 inhibitors as therapeutic agents. It is
clear that the D2 ATPase domain hydrolyzes ATP in vitro, but whether D1 contributes to ATPase activity is
NU
controversial. Here, we use Walker A and B mutants to demonstrate that D1 is capable of hydrolyzing ATP,
and show for the first time that nucleotide binding in the D2 domain increases the catalytic efficiency (kcat
MA
/Km) of D1 ATP hydrolysis 280-fold, by increasing kcat 7-fold and decreasing Km about 40-fold. We further
show that an ND1 construct lacking D2 but including the linker between D1 and D2 is catalytically active,
D
resolving a conflict in the literature. Applying enzymatic observations to small-molecule inhibitors, we show
TE
that four p97 inhibitors (DBeQ, ML240, ML241, and NMS-873) have differential responses to Walker A and
AC
CE
P
B mutations, to disease-causing IBMPFD mutations, and to the presence of the N-domain binding cofactor
protein p47. These differential effects provide the first evidence that p97 cofactors and disease mutations can
alter p97 inhibitor potency and suggest the possibility of developing context-dependent inhibitors of p97.
2
ACCEPTED MANUSCRIPT
Introduction
The highly conserved and abundant p97 protein belongs to the type II AAA ATPase enzyme family
and converts chemical energy from ATP hydrolysis into mechanical energy to unfold proteins or disassemble
PT
protein complexes. p97 plays a critical role in cellular processes such as Golgi membrane reassembly (1),
RI
membrane transport (2,3), degradation of proteins by the ubiquitin–proteasome system (UPS) (4,5),
SC
regulation of myofibril assembly (6), cell division (7), and protein aggregation (8-10). Because p97 functions
in many protein homeostatic regulatory processes, it is a potential therapeutic target for cancer and
NU
neurodegenerative diseases (11-13).
Each monomer of p97 ATPase is composed of three domains (the N, D1 ATPase, and D2 ATPase
MA
domains) joined by linker regions. D1 and D2 each form a hexameric ring, which stack to form a cylindrical
structure, as revealed by X-ray crystallography (14,15). Although it is clear that the D2 domain hydrolyzes
D
ATP in vitro, the level of D1-specific ATPase activity varies among different reports (16-18). Furthermore,
TE
genetic studies in yeast reporting D1 ATPase activity as an essential function are controversial. For example,
AC
CE
P
some mutants that inhibit ATP hydrolysis in the D1 domain affect growth (17), whereas others do not (19).
However, a consistent observation is that the nucleotide-binding state of D1 is tightly controlled (15,18).
Thus, it is important to clarify the role of the D1 domain in vitro, providing hypotheses that can then be
tested in cells.
Adjacent to the D1 domain is the N domain, which functions as a protein-protein interaction site that
binds p97 cofactors (20,21). By recruiting cofactor proteins, the N domain links the mechanochemical
ATPase activity of p97 to many substrate proteins, accounting in part for the broad range of cellular p97
functions. For example, cofactor p47 is required for p97-mediated membrane fusion (22), by mediating
binding of p97 to syntaxin 5 during the reassembly of Golgi stacks at the end of mitosis (23). Binding of p47
to the N-terminus alters the structure and function of p97 by significantly reducing the diameter of the p97
ring (22) and inhibiting ATPase activity (24). Hence, p47 supports communication between the N, D1 and
D2 domains of p97.
There are several additional lines of evidence for conformational crosstalk among the three domains
3
ACCEPTED MANUSCRIPT
of p97. Conformational changes occur during the ATPase cycle, including rotation of the D1 and D2 rings
with respect to each other, changes in the hexameric ring diameter, and positioning of the N-domain (25-29).
Additionally, pathogenic mutations in the N and D1 domains alter nucleotide binding to the D1 domain and
PT
ATPase activity in the D2 domain (29-31). Defining interactions between the D1 and D2 domains could
RI
provide opportunities for developing small-molecule inhibitors with diverse mechanisms of inhibition and
SC
perhaps diverse cellular effects.
In an effort to better understand the mechanistic and cellular functions of p97, a few potent and
NU
specific small molecule inhibitors of p97 ATPase activity have been identified via high-throughput
screening, followed by chemical optimization (32-35). Published reversible p97 ATPase inhibitors fall into
MA
two classes – ATP-competitive (DBeQ, ML240, ML241) and noncompetitive (NMS-873) inhibitors. NMS873 is thought to inhibit p97 by binding at the D1-D2 interdomain linker and stabilizing the D2-ADP-bound
D
form, thus breaking the catalytic cycle. Both competitive and noncompetitive inhibitors affect multiple p97-
TE
dependent processes in cells, including ubiquitin fusion degradation (UFD), endoplasmic reticulum
AC
CE
P
associated degradation (ERAD), autophagy, and cancer cell growth. However, investigations have not yet
shown whether competitive compounds inhibit ATPase activity arising from only D2 or from both D1 and
D2, and whether inhibitors are influenced by nucleotide binding states, disease mutations, or cofactor
binding.
In this study, we enzymatically characterized a set of p97 constructs and mutations to elucidate the
effect of small-molecule p97 inhibitors on D1 and D2 domain ATPase activities. We first clarified the in
vitro activity of the D1 domain. We found that an ND1-linker truncation is a robust, catalytically competent
ATPase whose activity is similar to that of the D1 domain in full-length p97. Interestingly, D1 showed
ATPase activity in the full-length protein only when the D2 domain was capable of binding nucleotides.
Using our enzymatically characterized set of p97 proteins, we addressed the domain specificity of four potent
p97 inhibitors on wild type (WT) p97. We then evaluated whether inhibition was sensitive to pathogenic
mutations in the ND1 domain or to the presence of bound p47, a p97 binding protein. DBeQ and NMS-873
inhibited both ATPase domains, whereas ML240 and ML241 were specific for D2. In addition, inhibition of
4
ACCEPTED MANUSCRIPT
D2 by ML240 and ML241 was altered by a pathogenic mutation in ND1 and upon p47 binding, indicating
domain communication within p97. Together, our results provide the framework for developing domain,
cofactor-complex, and pathway specific inhibitors (32), with the ultimate goal of validating p97 as a potential
PT
therapeutic target.
RI
Results
SC
The Human p97 D1 Domain is a Competent ATPase
To resolve the controversy over whether the isolated D1 domain can hydrolyze ATP in vitro, we
NU
prepared two truncated constructs: ND1 (residues 1-458) without the D1-D2 linker region, and ND1L (1480), which included the linker region (Fig. 1A). ND1 was used previously to conclude that D1 was inactive
MA
(16), whereas ND1L has been used in crystallography studies (31) and was recently reported to be
catalytically active (29). Specific activities with 600 M ATP at room temperature for human and mouse
TE
D
p97 were 5.20.3 and 5.10.2 nmol Pi/min/nmol p97, respectively. We found that the ATPase activity of
ND1 was 1.4% of the WT level, whereas ND1L had 79% of the WT activity (at 600 uM ATP; Fig. 1B).
AC
CE
P
Thus, ND1L was 56-fold more active than ND1. As a comparison to ND1L, we also prepared a truncation
containing the D1-D2 linker region plus the D2 domain (LD2; amino acids 461-806). LD2 existed in
dynamic multimer-monomer equilibrium (Figs. S1), and thus it was not surprising that its ATPase activity
was only 4% of WT values (Fig. 1B).
To address the difference in activities between ND1 and ND1L, we evaluated their stabilities by
differential scanning fluorimetry (DSF; Figs. 1C and D) and nucleotide binding by surface plasmon
resonance (SPR; Figs. 2A and B). Thermal stability studies showed that addition of the linker strongly
stabilized ND1 by 13 oC, to a temperature that approaches the first melting transition of full-length p97.
Additionally, ND1 showed multiple melting transitions, whereas ND1L gave a single, highly cooperative
melting curve. The increased thermal stability and single-state melting curve suggested that ND1L formed a
more stable protein than ND1. To determine whether ND1 bound weakly to nucleotides, leading to low
catalytic activity, we monitored binding of nucleotides to ND1 and ND1L by SPR at 20 oC, well below the
melting temperature for ND1. Both ND1 and ND1L bound to ATP, ATPS, or ADP with similar binding
5
ACCEPTED MANUSCRIPT
affinities and kinetics (within 2-fold) (Fig. S2). Hence, the linker region had only a small effect on
nucleotide-binding affinity, which did not account for the 56-fold increase in ATPase activity. Rather, we
concluded that the linker region rendered the D1 domain a competent ATP hydrolase mainly by substantially
PT
increasing its thermal stability. We cannot rule out that the linker also favorably altered the conformational
RI
dynamics of the protein.
SC
By directly comparing the two constructs that were used in the previous studies, we identified the
source of the conflicting results on D1 activity. We demonstrated that ND1L can hydrolyze ATP, which is in
NU
a good agreement with the recent report by Tang and coworkers (29). In contrast, ND1 cannot hydrolyze
ATP, which is consistent with the earlier report by Song and coworkers (16).
MA
Nucleotide Binding and Enzyme Kinetic Analysis of Walker A and B Mutants
We next directly compared the nucleotide binding affinities and activity of ND1L to full-length WT
D
p97 and active-site mutants. Walker A mutations (D1-K251A and D2-K524A) disrupt ATP and ADP
TE
binding, whereas Walker B mutations (D1-E305Q and D2-E578Q) disrupt activation of hydrolysis but allow
AC
CE
P
ATP and ADP binding (16,17,36) (Fig. 1A). For nucleotide binding analysis, full-length WT p97 and Walker
A mutants were assayed by SPR to compare the binding affinities of the D1 domain in the context of the fulllength p97 protein versus isolated ND1 domains (Figs. 2C, S2, and Table S2). Nucleotides bound to WT p97
with two distinct KD values. The tight binding affinity (132, 131, and 36.7 nM for ADP, ATP, ATPgS,
respectively) corresponded to the KD values obtained for ND1, ND1L, and the D2-K524A mutant (which is
unable to bind nucleotides in the D2 domain). Thus, the D1 site bound to nucleotides with KD values in the
100 nM range. Based on these results and the binding to the D1-K251A mutant (which is unable to bind
nucleotides in the D1 domain), we ascribed the second, weaker binding affinity to the D2 domain (3.9, 5.5,
and 1.1 M for ADP, ATP, ATPgS, respectively). It is noteworthy that analysis of nucleotide binding by
SPR resulted in KD values similar to those obtained by fluorescence polarization (see below) and tryptophan
fluorescence (18,31), but gave ~10-fold tighter binding affinities compared to values from ITC (18,31).
Overall, our SPR results supported the previous studies showing that D1 bound nucleotides more tightly than
did D2 (18,35).
6
ACCEPTED MANUSCRIPT
We next measured the activity of the Walker A and B mutations, to evaluate how nucleotide binding
or ATP hydrolysis in one domain may alter the ATPase activity of the other domain. We found that Walker
A mutations lowered ATPase activity more than did the Walker B mutations, supporting the importance of
PT
nucleotide binding in one domain for ATPase activity of the other domain (Fig. 3A, black font indicates the
RI
active domain in each protein). We included 0.01% Triton X-100 in our standard reaction buffer and
observed a 1.7 fold increase in ATPase activity for WT p97 upon addition of Triton X-100 (Fig. 3A). In
SC
general, the increase in ATPase activity by Triton X-100 was greater for the D1-active Walker B mutant
NU
(D2-E578Q) compared to the D2-active Walker B mutant (D1-E305Q) (Fig. 3A).
These activity data prompted us to measure steady-state kinetic constants (kcat, Km, and kcat/Km), in
MA
order to understand the enzymology of D1 ATPase activity in the context of full-length p97 and to evaluate
the crosstalk between the D1 and D2 domains (Table S3 and Figs. 3B and C). Published kinetic studies have
D
focused on only full-length WT p97 (16) and the full-length D1-E305Q mutant (37). Our data for these
TE
constructs were consistent with the published values (16,38). The most striking deleterious effects of Walker
AC
CE
P
mutations were the 22-fold reduction in kcat (from 7.5 to 0.29 min-1) (Fig. 3D) and the10-fold reduction in
catalytic efficiency (kcat/Km; from 0.026 to 0.0013 min-1uM-1) (Fig. 3F) for the Walker A mutation in the D2
domain (D2-K524A), compared to WT. Thus, the strongest effect on catalysis came from blocking binding
of nucleotides to the D2 domain. Simply blocking catalysis of D2 without blocking nucleotide binding (D2E578Q) gave a modest 3-fold effect on kcat and actually increased catalytic efficiency through a reduction in
Km (kcat/Km from 0.026 to 0.39 min-1uM-1). D1-E305Q also showed a slightly higher kcat/Km than WT. Taken
with D2-E578Q, the data suggested a cross-inhibitory effect on the activity of the D1 and D2 domains. These
observations were consistent with the negative cooperativity between the D1 and D2 rings shown for mouse
p97 (28) and with the recent demonstration that the presence of D2 inhibited D1 activity (29). Finally, it is
noteworthy that ND1L gave the same kinetic constants as D2-E578Q (Table S3 and Fig. 3C). Thus, ND1L
can serve as an accurate kinetic model of the D1 domain in the full-length p97.
It is noteworthy that adding detergent (Triton X-100) to the assay buffer was critical for observing
robust ND1L activity (Fig. 3C). Triton X-100 increased kcat 4.7-fold, and had no effect on K m. This result
7
ACCEPTED MANUSCRIPT
suggested that detergent facilitated catalysis, perhaps by aiding the release of prebound ADP from the
hexamer (18,29). It was also likely that detergent prevented protein aggregation and reduced protein binding
to the surface of the plastic assay plates. We have included 0.01% Triton X-100 in our standard reaction
PT
buffer (32,34) and detergents like Triton X-100 and Tween-20 are frequently added when testing small-
RI
molecule inhibitors to prevent compounds from forming colloids that can cause non-specific enzyme
SC
inhibition (39).
ATP-Competitive p97 Inhibitors, ML240 and ML241 are Selective for the D2 domain
NU
We previously identified three potent, ATP-competitive p97 inhibitors (DBeQ, ML240, and ML241;
Fig. 4A) (32,34). Nerviano Medical Sciences and Genentech, Inc., identified NMS-873 as a potent non-ATP-
MA
competitive p97 inhibitor, with potencies in the low nanomolar range (Table S4) (33,35). To investigate
whether these four p97 compounds inhibit the ATPase activity of D1, D2, or both domains, IC50 values were
D
determined for WT p97 and five variants in the presence of 200 M or 875 M ATP. Lower potencies were
TE
observed for inhibition of WT p97 (roughly 2 to 5-fold increases in IC50; Fig. 4 and Tables S4 and S5).
AC
CE
P
The ATP competitive inhibitors were sensitive to Walker A and B mutations in both domains. There
was a 2.7- to 10-fold increase in IC50 for D1-E305Q (Walker B mutant in the D1 domain) for DBeQ, ML240,
and ML241 (Figs. 4 B-D). These increases in IC50 can be partly explained by the 2.4-fold decrease in Km for
D1-E305Q, because ATP competitive inhibitors are sensitive to the levels of ATP binding (Fig. 3E). In
contrast to the smaller effects found for the D1 D1-E305Q mutant, ML240 was inactive toward D2 mutants
D2-E578Q and D2-K524A, and toward ND1L. ML241 was also inactive towards D2-K524A and ND1L, and
showed 147-fold weaker inhibition for D2-E578Q. Thus, both ML240 and ML241 were D2 selective
inhibitors (Figs. 4 C and D). DBeQ was less selective, showing 8 to 10-fold weaker inhibition of D2-E578Q,
D2-K524A, and ND1L (Fig. 4B). The “D2 specific” inhibitors were nevertheless sensitive to changes in D1.
For example, the D1 D1-K251A mutation increased the IC50 for ML240 and ML241 by 27-fold with 200 M
ATP and ML240 can not inhibit D1 D1-K251A with 875 M ATP, suggesting that the architecture of the D2
active site was changed significantly due to the mutation and/or due to the absence of nucleotide binding to
8
ACCEPTED MANUSCRIPT
D1. This result was consistent with our kinetic analysis of D1-K251A (Fig. 3B), which also showed strong
effects of this mutation on the activity of the D2 domain.
The allosteric inhibitor NMS-873 sensed mutations in the D2 domain, but was insensitive to D1
PT
mutations (Tables S4 and S5, Fig. 4E). In agreement with the published data for Walker A mutants, NMS-
RI
873 inhibited D1-K251A and WT p97 similarly, but could not bind and inhibit D2-K524A (35).
SC
Unexpectedly, NMS-873 inhibited D2-E578Q (IC50 1.0 M, or 25-fold higher than for WT p97, Fig. 4E),
suggesting that maintaining D2 in the ATP-bound state allowed NMS-873 to bind and prevent D1 from
NU
hydrolyzing ATP. In addition to the proposed mechanism, whereby NMS-873 prevented ADP release from
D2, our D2-E578Q data suggested that NMS-873 locked p97 into a conformation that may have also
MA
prevented ADP release from D1. The ATP-binding sites in D1 and D2 are similar for DBeQ binding, because
the IC50 values for D1-E305Q, D1-K251A, D2-E578Q, D2-K524A, and the isolated ND1L are almost
D
identical at higher concentrations of ATP (Table S5). Therefore, DBeQ inhibits both D1 and D2 ATPase
TE
activities similarly. However, ML240 and ML241 behave quite differently. They are much more sensitive to
AC
CE
P
Walker A and B mutations and cannot inhibit D2-E578Q, D2-K524A, and ND1L as potently as seen for D1E305Q. Thus, they are more selective for inhibiting D2 ATPase activity. Even though NMS-873 does not
bind to ATP binding sties, its ability to bind to p97 and inhibit D1 ATPase is diminished for the D2-K524A
mutant (35). However, when D1 and D2 can both bind nucleotides (as in D2-E578Q), NMS-873 inhibits D1
ATP hydrolysis, but 25-fold less potent as compared to D1-E305Q p97. In summary, the four inhibitors fall
into three classes: NMS-873, which binds at the D1-D2 interface, seems to inhibit ATPase activity from both
D1 and D2 domains; the ATP-competitive inhibitors DBeQ can inhibit both D1 and D2 ATPase activity; and
ML240 and ML241 are selective against the D2 domain.
Taken together, the D1 ATPase activity is more difficult to block by p97 inhibitors (Figs. 4 B to E,
D2-E578Q). This is consistent with results of our steady state kinetic analysis of D2-E578Q, the mutant with
the lowest Km for ATP (Figs. 3B, C, and E) and tighter nucleotide binding to D1 (Table S2 and Figs. 2 and
5A). Furthermore, all ATP-competitive inhibitors (DBeQ, ML240, and ML241) displayed lower inhibitory
potencies (higher IC50 values) toward all ATPase-defective mutants (Walker A/B mutants), compared to the
9
ACCEPTED MANUSCRIPT
WT. On the other hand, the non-ATP-competitive inhibitor NMS-873 showed similar potencies toward the
D1 mutants (D1-E305Q and D1-K251A), compared to the WT, but much lower potencies toward D2 mutants
(D2-E578Q and D2-K524A). In other words, the mutants in general are more resistant to inhibitors,
PT
especially the D2 Walker A mutant (D2-K524A). Such resistance seems expected, due to the significant
RI
impairment of their steady state behavior toward ATP.
SC
To determine if D2 domain-selective inhibitors affect ATP binding to the D1 domain, we performed
experiments with fluorescent BODIPY-FL-ATP. Fluorescence anisotropy was detected when BODIPY-FL-
NU
ATP bound to WT p97 (Fig. 5A, red line). The KD values for BODIPY-FL-ATP binding were similar for WT
p97 and the Walker A mutation in the D2 domain (D2-K524A, Fig. 5A, blue line), indicating that nucleotide
MA
binding monitored in WT p97 was predominantly to D1. As anticipated by SPR data, BODIPY-FL-ATP
does bind to D2 (D1-K251A, Figure 5A, black line) with a ~23-fold weaker affinity than for D1. These
D
results are also consistent with the binding of other fluorescent probes to WT p97 and Walker A mutants
TE
(18,35). We then carried out binding experiments in the presence of 50 M ML240 and ML241 and
(Fig. 5B).
AC
CE
P
demonstrated that inhibiting D2 with ML240 and ML241 did not affect BODIPY-FL-ATP binding to D1
IBMPFD/ALS Mutations and p47 Cofactor Binding Affect Potency of the D2-selective p97 Inhibitors
Mutations at the ND1 interface of p97 cause IBMPFD/ALS (40,41), a devastating degenerative
disorder that has no known treatment. Therefore, p97 inhibitors that can correct the increased ATPase
activity of mutants to WT levels might be useful for exploration in animal models, as a possible starting point
for research on treatment approaches.
To begin to investigate this opportunity, we prepared a series of IBMPFD p97 mutants and tested the
activity of ML240 and ML241. We first confirmed a prior report (38) that the elevated ATPase activity from
mutants such as A232E arose from turnover in D2 (Fig. 6A). As expected, when D1 activity was knocked out
by D1-E305Q, D2 activity was increased by A232E, whereas A232E had no affect on D1-mediated ATPase
activity (D2-E578Q). We then determined IC50 values for the D2-selective inhibitors, ML240 and ML241,
against eight mutants that cause disease (Fig. 6B). Interestingly, some mutants were less sensitive to the
10
ACCEPTED MANUSCRIPT
inhibitors. For example, the IC50 value for ML240 was 5-fold higher for the R95G mutant, compared to WT
p97. This result demonstrated that the conformation at the ND1 interface altered binding of inhibitors to the
D2 ATP site. Our data suggest that it may be possible to develop a mutant selective inhibitor. Selective
PT
inhibition of the mutant subunit may reduce the pool of active p97 mutants (or hexamers containing mutants)
RI
within cells, so that cells use predominantly WT hexamers. However, additional work will be required to
SC
develop mutant-selective inhibitors.
Inhibitor potency was clearly affected by IBMPFD/ALS mutations and by the nucleotide binding
NU
state of the D1 domain of p97, even though these mutations are distal to the small-molecule binding site in
the D2 domain. By extension, we hypothesized that cofactor binding to p97 might also be critical for p97
MA
inhibitor potency. Therefore, we evaluated inhibitor potency for p97–p47 complexes at four concentrations
of p47 (Table S6 and Fig. 7). All concentrations of p47 increased the IC50 of DBeQ 4- to 6-fold (Fig. 7A).
D
More dramatically, p47 trimer concentrations greater than 16.7 nM increased IC50 values of ML240 and
TE
ML241 more than 49-fold and 37-fold, respectively (Figs. 7B and C). In contrast, only a minor 2-fold effect
AC
CE
P
was detected for the allosteric inhibitor NMS-873 (Fig. 7D). This result provided the first demonstration that
a major p97 cofactor, p47, can influence the potency of ATP-competitive p97 inhibitors.
Discussion
Our interest is to understand the function of the D1 and D2 p97 ATPase domains, through kinetic
analysis of mutants, and use of small molecule inhibitors. The impetus to evaluate effects of p97 inhibitors
came from work on the D1 inhibitor, Eeyarestatin I (27), and the N-domain inhibitor, Xanthohumo1 (28).
Neither of these compounds blocks full-length p97 ATPase activity, but they inhibit some p97-dependent
cellular activities. The results suggest that modifying activity of either the N or D1 domain by small
molecules may alter different p97 functions in cells. Moreover, the human disorder known as IBMPFD/ALS
is associated with dominantly inherited, missense mutations in the N, ND1 linker, and D1 domains (that
cause high p97 ATPase activity), but not by mutations in D2 (29,30).
The AAA ATPase p97 is a dynamic molecular machine that consists of two ATPase domains, D1
and D2, which function to drive the molecular remodeling of protein substrates. We first did a systematic
11
ACCEPTED MANUSCRIPT
comparison of D1 and D2 ATPase activity to resolve the controversy on whether D1 is an active ATPase.
With the active ND1L mutant and Walker A and B mutants for both D1 and D2, we carried out steady-state
kinetic analyses and dose-dependent titrations with 4 p97 inhibitors against these mutants.
PT
Our steady-state kinetic analyses provide molecular insights into the roles of both ATPase domains,
RI
and also indicate that nucleotide binding and p97 inhibitors modulate each domain interdependently. We
show that the ND1L domain is a catalytically competent ATPase, resolving a debate in the literature. Tang
SC
and coworkers (29) demonstrated ATPase activity in an ND1L construct (Fig. 1A), whereas two other labs
NU
published that ND1 was inactive (16,18). Moreover, our results provide a plausible explanation for the
conflicting reports on whether D1 is an active ATPase. We show that the D1-D2 linker region (residues 459-
MA
480) dramatically improves thermal stability (Fig. 1D). However, both ND1 and ND1L are capable of
binding nucleotides when the temperature is below the melting temperature (Figs. 2A and B). These data are
D
consistent with X-ray crystallography studies that reveal identical active-site structural arrangements for
TE
ND1, ND1L, and the ND1 domain in the full-length enzyme (14,15,31,42). One possible explanation for the
56-fold increase in activity in the ND1L construct is that the linker region positions a water molecule into a
AC
CE
P
catalytically competent orientation to hydrolyze ATP.
In this study, we provide quantitative analysis of kinetic constants for the D1 and D2 ATPase
domains of human p97, by using point mutations at the conserved Walker A and B positions of D1 and D2
(16,17,36) and an isolated ND1L domain. Additionally, we examine the binding kinetics and dissociation
constants for Walker A mutations and isolated ND1 domains. Our estimated Km of 287 M for ATP at room
temperature for human p97 is within the range of published values. Specifically, estimated K m values for p97
isolated from rat liver are 550 M or 620 M (24,43) for ATP at 37 oC. Three studies found K m values of
330 M (16), 360 M (38), or 49 M (44) for recombinant mouse p97 purified from E. coli. The Km for the
C. elegans p97 homolog is 390 M (37). Some variation among different reports may be due to differences
in assay methods.
Our results showed that ATP hydrolysis within the D1 domain of a full-length p97 protein containing
a Walker B mutation in D2 (D2-E578Q) exhibited Michaelis-Menten behavior almost identical to that for
12
ACCEPTED MANUSCRIPT
ND1L (Fig. 3C, Table S3). Consistent with the enzymology, isolated ND1L was found to bind nucleotides
with similar affinities as the D1 domain in full-length p97 (Figs. 2 and S2). Furthermore, the Km of D1 (D2E578Q) for ATP was ~20-fold less than the D2 value (D1-E305Q) (Fig. 3E), consistent with the 10-fold
PT
tighter nucleotide binding to D1 than to D2 (Table S2) (18). Although the D1 maximal velocity was about
RI
3.5 times lower than for WT p97 (Figs. 3C and D), the kcat/Km for D1 is actually 14-fold higher than the WT
SC
value, indicating that D1 is a catalytically competent ATPase within full-length p97.
Previous studies hinted that intra- and inter-domain communication within the p97 hexamer is
NU
important for p97 activity (17,27-29,37). We found that preventing nucleotide binding in the D2 domain
using a Walker A mutant (D2-K524A) drastically impaired both the D1 Km for ATP (~40-fold weaker) and
MA
catalytic activity (7-fold lower kcat), as compared to a D2-E578Q mutant (Fig. 3A). This striking finding
indicates that nucleotide binding to D2 is required for D1 activity in full-length p97. Hence, domain
D
communication within the p97 hexamer critically controls the capacity of both domains to hydrolyze ATP.
TE
Moreover, the tighter Km for D1 may imply that D1 ATPase activity is more important under stress
AC
CE
P
conditions, when cellular ATP concentration is depleted (45-48).
In contrast to our experiments with ND1L, a truncation strategy was not suitable for investigating the
catalytic efficiency of the D2 ATPase, because isolated D2 cannot form stable hexamers (Fig. S1) (49),
Instead, we focused on D2 ATP hydrolysis within the full-length p97 protein. The Walker B mutation at D1
(D1-E305Q) resulted in a 2.4-fold decrease in Km for the D2 domain, which increased its catalytic efficiency
(kcat/Km) by ~ 3-fold (Table S3). Conversely, the Walker A mutation in D1 (D1-K251A) prevented D2 from
reaching maximal WT activity, no matter how much ATP was present. Because the Km values are similar for
D1-D1-K251A and D1-E305Q (Figs. 3A and D), we conclude that nucleotide binding to D1 is essential to
achieve maximal kcat values for D2 (18,37).
Together, our results support the presence of domain communication within the full-length WT p97
ATPase, in which nucleotide binding to one ATPase domain is essential for activity of the other domain.
Further work is needed to decipher the precise roles and the molecular mechanisms of this communication.
In particular, the importance of D1 ATPase activity in cells is controversial (17,19,29), and our enzymatic
13
ACCEPTED MANUSCRIPT
studies with mutants pave the way to investigate the cellular role of D1 activity, cooperativity between D1
and D2, and the extent to which p97 cofactors are involved in regulating such communication.
Our results also have implications for the design of p97 inhibitors for use in cancer and
PT
neurodegeneration. We have generated a set of tools to develop D1- and D2-domain specific inhibitors, and
RI
clearly demonstrated the existence of two types of ATP-competitive inhibitors. DBeQ inhibits both D1 and
SC
D2, whereas ML240 and ML241 only inhibit the D2 ATPase domain (Figs. 4 A-C and Tables S4 and S5).
Moreover, inhibition by ML240 and ML241 is sensitive to IBMPFD/ALS mutations in p97 (Fig. 6B),
NU
suggesting that IBMPFD/ALS-selective inhibitors might be developed via either structure-based drug design
or high-throughput screens.
MA
Another significant finding in this study is the roughly 50-fold decrease in ML240 potency in the
presence of p47, providing the first evidence that p97 cofactors can affect the potency of ATP-competitive
D
inhibitors (Figs. 7A-C, and Table S6). These results favor an interpretation that the p97–p47 complex exists
TE
in a conformation that cannot be inhibited by ML240 and ML241 as effectively as by the allosteric inhibitor
AC
CE
P
NMS-873. Indirect evidence suggests that p97 ATP binding sites exist in different conformations in the
presence of p47, because p47 dramatically changes the global p97 conformation (21,22,50) and lowers the
maximal velocity of ATPase activity by 5-fold (24). Thus, it may be possible to develop complex-specific
p97 inhibitors for p47 and other p97 cofactor proteins (51). It is enticing to consider whether ML240 and
ML241 may inhibit a subset of p97 complexes in cells, whereas NMS-873 may be a broad-spectrum
inhibitor. Future in vivo studies are needed to compare anti-cancer potencies and general toxicities of distinct
classes of p97 inhibitors. Taken together, our results provide a new way to consider domain communication
in p97 and the tools to develop D1- or D2-domain selective, complex-specific, and disease-mutant-selective
p97 inhibitors.
Materials and Methods
Materials- Plasmids used are listed in Supplemental Table S1.
BIOMOL Green ATPase Assay— Purified p97 (25 L of 50 M; final concentration in the reaction was 25
nM) was diluted in 40 mL of assay buffer [10 mL of 5x assay buffer A (1x = 50 mM Tris pH 7.4, 20 mM
14
ACCEPTED MANUSCRIPT
MgCl2, 1 mM EDTA,) mixed with 30 mL water and 50 L 0.5M TCEP, 50 L 10% Triton] to make the
enzyme solution. Some experiments represented in Figures 3A and 3C were performed without 0.01%
Triton X-100. In these experiments, 50 L 10% Triton was not added to the assay buffer. For all other
PT
experiments, 50 L 10% Triton was included. 40 L of enzyme solution was dispensed into each well of a
RI
96 well plate. The WT p97 ATPase assay was carried out by adding 10 L of 1000 M or 4375 M ATP
SC
(Roche, pH 7.5) to each well and incubating the reaction at room temperature for 35 or 15 min, respectively
Reactions were stopped by adding 50 L of BIOMOL Green reagent (Enzo Life Sciences). Absorbance at
NU
635 nm was measured after 4 min on the Synergy Neo Microplate Reader (BioTek). To determine the
steady-state kinetic constant, eight final ATP concentrations were used (62.5, 125, 250, 375, 500, 625, 750,
MA
875 M) or (3.9, 7.8, 15.6, 23.4, 31.25, 62.5, 93.8, 125 M). For mutants, reaction times were adjusted
according to specific activities, to give acceptable absorbance readings. Michaelis-Menten constants were
TE
ν = kcat [E]t *[S]/(Km + [S])
D
calculated from eight replicates by fitting data to Equation 1 using GraphPad Prism 6.0.
(1)
AC
CE
P
where ν is the initial velocity, kcat is the turnover number, [E]t is the total enzyme concentration, [S] is the
substrate concentration, and Km is the concentration of substrate to reach half maximal velocity.
ATPase Assay for Determining IC50 Values of Inhibitors— Purified p97 (25 L of a 50 M stock solution)
was diluted in 30 mL of assay buffer [10 mL of 5 x assay buffer A (1 x = 50 mM Tris pH 7.4, 20 mM MgCl2,
1 mM EDTA,) mixed with 20 mL of water, 50 L of 0.5M TCEP, and 50 L of 10% Triton X-100] to make
the enzyme solution. 30 L of enzyme solution was dispensed into each well of a 96 well plate. The test
compound (10 µL) or DMSO (12%, 10 L) was then added to each well, and the plate was incubated at
room temperature for 10 min. The reaction was initiated by adding 10 L of 1 mM ATP (pH 7.5) to each
well at room temperature. After 30 min, the reaction was stopped by adding 50 L of BIOMOL Green
reagent. Absorbance at 635 nm was measured after 4 min. DBeQ was assayed at a range of concentrations (0,
0.04, 0.12, 0.37, 1.1, 3.3, 10, 30 M). The other compounds were assayed at (0, 0.013, 0.04, 0.12, 0.37, 1.1,
3.3, and 10 M). IC50 values were calculated from four replicates by non-linear regression analysis with
15
ACCEPTED MANUSCRIPT
Equation 2 using GraphPad Prism 6.0.
% Activity =100/(1+10^((X-LogIC50)))
(2)
PT
where X is the compound concentration.
Surface Plasmon Resonance (SPR)— The kinetic parameters for the direct binding of nucleotides to p97
RI
were measured on a Biacore 4000 instrument. Neutravidin-coated sensor chips were prepared on CM5 chips,
SC
as follows: the surface was activated by injecting a 1:1 mixture of 60 mM N-hydroxysuccinimide and 240
mM 1-ethyl-3-(3-dimethylaminopropyl)-carbodiimide for 7 min, followed by a 7 min injection of 0.25
NU
mg/mL NeutrAvidin (Thermo Scientific) in 10 mM acetic acid pH 4.5. The surface was then blocked by a 2
MA
min injection of 1 M ethanolamine, pH 8.3.
p97 proteins containing site-specific biotinylation (AviTag; Avidity) were expressed and purified
D
from E. coli (see Supplemental materials). Purified protein was immobilized in 10 mM HEPES pH 7.5, 150
TE
mM NaCl, 0.5 mM TCEP, 0.05% Tween 20. Protein was immobilized to 4000-6000 resonance units (RUs)
by injecting 100-500 g/mL of protein for 1 min at 25 oC. Protein immobilization was followed by a 2 min
AC
CE
P
injection of 0.2 mg/mL amino-PEG-biotin (Thermo Scientific) to block any remaining biotin binding sites. A
reference surface was created using the same protocol with omission of the protein injection step.
Binding of ADP (Sigma A2754), ATP
measured in duplicate in SPR running buffer (25 mM Tris pH 7.5, 150 mM NaCl, 10 mM MgCl 2, 0.5 mM
TCEP, 0.05% Tween 20) at 20 °C. Nucleotides were injected at a flow rate of 30 L/min for 90-180 seconds,
with a dissociation time of 180-600 seconds.
SPR experiments were carried out in the presence of 0.05% Tween-20, whereas the BIOMOL Green
ATPase assays were carried out in the presence of 0.01% Triton X-100. Because Tween-20 and the
BIOMOL Green reagent are incompatible, to compare the ATPase activities of WT p97 in the presence of
Triton X-100 and Tween-20, we used the ADP-Glo assay. Similar to the BIOMOL green assay, we found the
addition of 0.01% Triton X-100 or Tween-20 modestly increased WT p97 activity by 1.79- and 1.97-fold,
16
ACCEPTED MANUSCRIPT
respectively. The addition of 0.05% Tween-20 further increased the activity by 2.8-fold over no detergent
conditions.
Differential Scanning Fluorimetry (DSF) — The thermal stabilities of p97 ND1 and ND1L domains were
PT
measured on a Mx3005P q-PCR (Stratagene) instrument in a 24-well Thermo-Fast PCR plate (Thermo
RI
Scientific). Four replicates of 20 L of ND1, ND1L or full-length p97 (2.0 M) were mixed with 5 L of a
SC
1:200 dilution of SYPRO Orange Protein Gel Stain (Sigma) in 25 mM Tris, pH 7.5, 150 mM NaCl, 10 mM
MgCl2, 0.5 mM TCEP. Samples were heated from 25 °C to 100 °C at 1 °C/min, and fluorescence (Ex = 492
NU
nM, Em = 610 nM) was measured every minute. Curves were normalized, and the melting temperature was
calculated from the first derivative using the FIT for DSF website (skuld.bmsc.washington.edu/FIT/).
MA
Fluorescence Anisotropy — Fluorescence anisotropy experiments were completed in black, low volume,
non-binding 384-well plates (Corning) in 25 mM Tris, pH 7.5, with 150 mM NaCl, 10 mM MgCl2, and 0.5
D
mM TCEP, by mixing 10 L of protein (200 M – 48 pM; 2-fold dilutions) with 10 L of BODIPY-FL-ATP
TE
ligand (Life Technologies; 20 nM). After a 30-minute incubation, fluorescence polarization was measured
AC
CE
P
on an Analyst HT plate reader (Molecular Devices) using 485 nm and 530 nm excitation and emission filters,
respectively. No time-dependent change in anisotropy was observed over 30 minutes (e.g., due to catalytic
turnover of the BODIPY-FL-ATP probe). Background fluorescence (protein only) was subtracted from the
parallel and perpendicular intensities before calculating anisotropy. The KD of BODIPY-FL-ATP was
determined with GraphPad Prism 5.0, by Equations 3 and 4 to account for ligand depletion:
[RL] = ((KD + [L] + [R]) - √((KD + [L] + [R])2 – 4[L][R]))/2
(3)
A = Af + (Ab – Af ) x [RL]
(4)
where [RL] is the concentration of the p97-BODIPY-FL-ATP complex, [L] is the concentration of BIODPYFL-ADP, [R] is the concentration of p97, A is anisotropy, Af is the anisotropy of BODIPY-FL ATP, and Ab
is the anisotropy of the p97-BODIPY-FL-ATP complex.
To determine if DBeQ, ML240, or ML241 altered nucleotide binding to the D1 domain of p97, 5 L
of WT p97 (100 M – 24 pM; 2-fold dilutions) was mixed with 5 L of 5% DMSO or compound (200 M in
5% DMSO buffer) and 10 μL of 20 nM BODIPY-FL-ATP.
17
ACCEPTED MANUSCRIPT
Materials and additional methods, including procedures for protein expression, are available in
supplemental data.
Acknowledgment. We thank Michelina Iacovino for critical reading and suggestions on this manuscript. We
PT
thank David Myszka for assistance in fitting SPR data in CLAMP. The project was in part supported by the
RI
National Center for Advancing Translational Sciences through UCLA CTSI Grant UL1TR000124 and the
SC
LA BioMed Seed Grant program (20826-01). This project has been funded in part with federal funds from
the National Cancer Institute, National Institutes of Health, under Chemical Biology Consortium Contract
NU
No. HHSN261200800001E and by the National Human Genome Research Institute of the National Institutes
of Health under Award Number U54HG005031 (Jeffrey Aubé, PI). The content of this publication does not
MA
necessarily reflect the views or policies of the Department of Health and Human Services, nor does mention
of trade names, commercial products, or organizations imply endorsement by the U.S. government. RJD is
1.
2.
3.
4.
5.
6.
7.
8.
9.
TE
References:
AC
CE
P
Comprehensive Cancer Center.
D
an Investigator of the Howard Hughes Medical Institute, and TFC is a member of UCLA Jonsson
Rabouille, C., Levine, T. P., Peters, J. M., and Warren, G. (1995) An NSF-like ATPase, p97,
and NSF mediate cisternal regrowth from mitotic Golgi fragments. Cell 82, 905-914
Ye, Y., Meyer, H. H., and Rapoport, T. A. (2001) The AAA ATPase Cdc48/p97 and its partners
transport proteins from the ER into the cytosol. Nature 414, 652-656
Ye, Y., Shibata, Y., Yun, C., Ron, D., and Rapoport, T. A. (2004) A membrane protein complex
mediates retro-translocation from the ER lumen into the cytosol. Nature 429, 841-847
Golbik, R., Lupas, A. N., Koretke, K. K., Baumeister, W., and Peters, J. (1999) The Janus face
of the archaeal Cdc48/p97 homologue VAT: protein folding versus unfolding. Biol. Chem. 380,
1049-1062
Richly, H., Rape, M., Braun, S., Rumpf, S., Hoege, C., Jentsch, S. (2005) A series of ubiquitin
binding factors connects CDC48/p97 to substrate multiubiquitylation and proteasomal
targeting. Cell 120, 73-84
Janiesch, P. C., Kim, J., Mouysset, J., Barikbin, R., Lochmuller, H., Cassata, G., Krause, S.,
and Hoppe, T. (2007) The ubiquitin-selective chaperone CDC-48/p97 links myosin assembly to
human myopathy. Nat. Cell Biol. 9, 379-390
Cao, K., Nakajima, R., Meyer, H. H., and Zheng, Y. (2003) The AAA-ATPase Cdc48/p97
regulates spindle disassembly at the end of mitosis. Cell 115, 355-367
Boyault, C., Gilquin, B., Zhang, Y., Rybin, V., Garman, E., Meyer-Klaucke, W., Matthias, P.,
Muller, C. W., and Khochbin, S. (2006) HDAC6-p97/VCP controlled polyubiquitin chain
turnover. EMBO J. 25, 3357-3366
Boyault, C., Zhang, Y., Fritah, S., Caron, C., Gilquin, B., Kwon, S. H., Garrido, C., Yao, T. P.,
Vourc'h, C., Matthias, P., and Khochbin, S. (2007) HDAC6 controls major cell response
pathways to cytotoxic accumulation of protein aggregates. Genes Dev. 21, 2172-2181
18
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
NU
SC
RI
PT
10. Ju, J. S., Miller, S. E., Hanson, P. I., and Weihl, C. C. (2008) Impaired protein aggregate
handling and clearance underlie the pathogenesis of p97/VCP-associated disease. J. Biol.
Chem. 283, 30289-33099
11. Fessart, D., Marza, E., Taouji, S., Delom, F., and Chevet, E. (2013) P97/CDC-48: proteostasis
control in tumor cell biology. Cancer letters 337, 26-34
12. Stolz, A., Hilt, W., Buchberger, A., and Wolf, D. H. (2011) Cdc48: a power machine in protein
degradation. Trends Biochem. Sci. 36, 515-523
13. Kakizuka, A. (2008) Roles of VCP in human neurodegenerative disorders. Biochem.
Soc. Trans. 36, 105-108
14. Huyton, T., Pye, V. E., Briggs, L. C., Flynn, T. C., Beuron, F., Kondo, H., Ma, J., Zhang, X.,
and Freemont, P. S. (2003) The crystal structure of murine p97/VCP at 3.6A. J. Struct. Biol.
144, 337-348
15. DeLaBarre, B., and Brunger, A. T. (2003) Complete structure of p97/valosin-containing protein
reveals communication between nucleotide domains. Nat. Struct. Biol. 10, 856-863
16. Song, C., Wang, Q., and Li, C. C. (2003) ATPase activity of p97-valosin-containing protein
(VCP). D2 mediates the major enzyme activity, and D1 contributes to the heat-induced activity.
J. Biol. Chem. 278, 3648-3655
17. Ye, Y., Meyer, H. H., and Rapoport, T. A. (2003) Function of the p97-Ufd1-Npl4 complex in
retrotranslocation from the ER to the cytosol: dual recognition of nonubiquitinated polypeptide
segments and polyubiquitin chains. J. Cell Biol. 162, 71-84
18. Briggs, L. C., Baldwin, G. S., Miyata, N., Kondo, H., Zhang, X., and Freemont, P. S. (2008)
Analysis of nucleotide binding to p97 reveals the properties of a tandem AAA hexameric
ATPase. J. Biol. Chem. 283, 13745-11352
19. Esaki, M., and Ogura, T. (2010) ATP-bound form of the D1 AAA domain inhibits an essential
function of Cdc48p/p97. Biochem. Cell Biol. 88, 109-117
20. Dreveny, I., Kondo, H., Uchiyama, K., Shaw, A., Zhang, X., and Freemont, P. S. (2004)
Structural basis of the interaction between the AAA ATPase p97/VCP and its adaptor protein
p47. EMBO J. 23, 1030-1039
21. Uchiyama, K., Jokitalo, E., Kano, F., Murata, M., Zhang, X., Canas, B., Newman, R.,
Rabouille, C., Pappin, D., Freemont, P., and Kondo, H. (2002) VCIP135, a novel essential
factor for p97/p47-mediated membrane fusion, is required for Golgi and ER assembly in vivo.
J. Cell. Biol. 159, 855-866
22. Kondo, H., Rabouille, C., Newman, R., Levine, T. P., Pappin, D., Freemont, P., and Warren, G.
(1997) p47 is a cofactor for p97-mediated membrane fusion. Nature 388, 75-78
23. Rabouille, C., Kondo, H., Newman, R., Hui, N., Freemont, P., and Warren, G. (1998) Syntaxin
5 is a common component of the NSF- and p97-mediated reassembly pathways of Golgi
cisternae from mitotic Golgi fragments in vitro. Cell 92, 603-610
24. Meyer, H. H., Kondo, H., and Warren, G. (1998) The p47 co-factor regulates the ATPase
activity of the membrane fusion protein, p97. FEBS Lett 437, 255-257
25. Rouiller, I., DeLaBarre, B., May, A. P., Weis, W. I., Brunger, A. T., Milligan, R. A., and WilsonKubalek, E. M. (2002) Conformational changes of the multifunction p97 AAA ATPase during its
ATPase cycle. Nat. Struct. Biol. 9, 950-957
26. Noi, K., Yamamoto, D., Nishikori, S., Arita-Morioka, K., Kato, T., Ando, T., and Ogura, T.
(2013) High-speed atomic force microscopic observation of ATP-dependent rotation of the
AAA+ chaperone p97. Structure 21, 1992-2002
27. Davies, J. M., Tsuruta, H., May, A. P., and Weis, W. I. (2005) Conformational changes of p97
during nucleotide hydrolysis determined by small-angle X-Ray scattering. Structure 13, 183195
28. Beuron, F., Flynn, T. C., Ma, J., Kondo, H., Zhang, X., and Freemont, P. S. (2003) Motions
and negative cooperativity between p97 domains revealed by cryo-electron microscopy and
quantised elastic deformational model. J. Mol. Biol. 327, 619-629
29. Tang, W. K., and Xia, D. (2013) Altered Intersubunit Communication Is the Molecular Basis for
Functional Defects of Pathogenic p97 Mutants. J. Biol. Chem. 288, 36624-33635
19
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
NU
SC
RI
PT
30. Halawani, D., LeBlanc, A. C., Rouiller, I., Michnick, S. W., Servant, M. J., and Latterich, M.
(2009) Hereditary inclusion body myopathy-linked p97/VCP mutations in the NH2 domain and
the D1 ring modulate p97/VCP ATPase activity and D2 ring conformation. Mol. Cell Biol. 29,
4484-4494
31. Tang, W. K., Li, D., Li, C. C., Esser, L., Dai, R., Guo, L., and Xia, D. (2010) A novel ATPdependent conformation in p97 N-D1 fragment revealed by crystal structures of diseaserelated mutants. EMBO J. 29, 2217-2229
32. Chou, T.-F., Brown, S. J., Minond, D., Nordin, B.E., Li, K., Jones, A.C., Chase, P., Porubsky,
P. R., Stoltz, B.M., Schoenen, F. J., Patricelli, M.P., Hodder, P., Rosen, H., and Deshaies, R.
J. (2011) Reversible inhibitor of p97, DBeQ, impairs both ubiquitin-dependent and autophagic
protein clearance pathways. Proc. Natl. Acad. Sci. USA 108, 4834-4839
33. Polucci, P. e. a., and D'Alessio, R. (2013) Alkylsulfanyl-1,2,4-triazoles, a new class of
allosteric valosine containing protein inhibitors. Synthesis and structure-activity relationships.
J. Med. Chem. 56, 437-450
34. Chou, T. F., Li, K., Frankowski, K. J., Schoenen, F. J., and Deshaies, R. J. (2013) Structureactivity relationship study reveals ML240 and ML241 as potent and selective inhibitors of p97
ATPase. ChemMedChem 8, 297-312
35. Magnaghi, P., et. al. and Isacchi, A. (2013) Covalent and allosteric inhibitors of the ATPase
VCP/p97 induce cancer cell death. Nat. Chem. Biol. 9, 548-556
36. Walker, J. E., Saraste, M., Runswick, M. J., and Gay, N. J. (1982) Distantly related sequences
in the alpha- and beta-subunits of ATP synthase, myosin, kinases and other ATP-requiring
enzymes and a common nucleotide binding fold. EMBO J 1, 945-951
37. Nishikori, S., Esaki, M., Yamanaka, K., Sugimoto, S., and Ogura, T. (2011) Positive
cooperativity of the p97 AAA ATPase is critical for essential functions. J. Biol. Chem. 286,
15815-15820
38. Niwa, H., Ewens, C. A., Tsang, C., Yeung, H. O., Zhang, X., and Freemont, P. S. (2012) The
role of the N-domain in the ATPase activity of the mammalian AAA ATPase p97/VCP. J. Biol.
Chem. 287, 8561-8570
39. McGovern, S. L., Caselli, E., Grigorieff, N., and Shoichet, B. K. (2002) A common mechanism
underlying promiscuous inhibitors from virtual and high-throughput screening. J. Med. Chem.
45, 1712-1722
40. Watts, G. D., Wymer, J., Kovach, M. J., Mehta, S. G., Mumm, S., Darvish, D., Pestronk, A.,
Whyte, M. P., and Kimonis, V. E. (2004) Inclusion body myopathy associated with Paget
disease of bone and frontotemporal dementia is caused by mutant valosin-containing protein.
Nat. Genet. 36, 377-381
41. Weihl, C. C., Miller, S. E., Hanson, P. I., and Pestronk, A. (2007) Transgenic expression of
inclusion body myopathy associated mutant p97/VCP causes weakness and ubiquitinated
protein inclusions in mice. Hum. Mol. Genet. 16, 919-928
42. Zhang, X., et al. and Freemont, P. S. (2000) Structure of the AAA ATPase p97. Mol. Cell 6,
1473-1484
43. Zhang, L., Ashendel, C. L., Becker, G. W., and Morre, D. J. (1994) Isolation and
characterization of the principal ATPase associated with transitional endoplasmic reticulum of
rat liver. J. Cell Biol. 127, 1871-1883
44. DeLaBarre, B., Christianson, J. C., Kopito, R. R., and Brunger, A. T. (2006) Central pore
residues mediate the p97/VCP activity required for ERAD. Mol. Cell 22, 451-462
45. Imamura, H., Nhat, K. P., Togawa, H., Saito, K., Iino, R., Kato-Yamada, Y., Nagai, T., and Noji,
H. (2009) Visualization of ATP levels inside single living cells with fluorescence resonance
energy transfer-based genetically encoded indicators. Proc. Natl. Acad. Sci. USA 106, 1565115656
46. Leppanen, O., Bjornheden, T., Evaldsson, M., Boren, J., Wiklund, O., and Levin, M. (2006)
ATP depletion in macrophages in the core of advanced rabbit atherosclerotic plaques in vivo.
Atherosclerosis 188, 323-330
20
ACCEPTED MANUSCRIPT
SC
RI
PT
47. St-Pierre, J., Brand, M. D., and Boutilier, R. G. (2000) Mitochondria as ATP consumers:
cellular treason in anoxia. Proc. Natl. Acad. Sci. USA 97, 8670-8674
48. Chandel, N. S., and Budinger, G. R. (2007) The cellular basis for diverse responses to oxygen.
Free Radic Biol Med 42, 165-174
49. Wang, Q., Song, C., and Li, C. C. (2003) Hexamerization of p97-VCP is promoted by ATP
binding to the D1 domain and required for ATPase and biological activities. Biochem. Biophys.
Res. Commun. 300, 253-260
50. Rouiller, I., Butel, V. M., Latterich, M., Milligan, R. A., and Wilson-Kubalek, E. M. (2000) A
major conformational change in p97 AAA ATPase upon ATP binding. Mol Cell 6, 1485-1490
51. Chou, T. F., and Deshaies, R. J. (2011) Development of p97 AAA ATPase inhibitors.
Autophagy 7, 1091-1092
NU
FIGURE LEGENDS
Figure 1. Comparison of WT and isolated domain variants. (A) Human p97 mutants and
MA
variants generated in this study. (B) ATPase activities of WT p97, ND1, ND1L, and LD2 domains
with 0.01% Triton X-100 at 600 M ATP. (C, D) The linker region (a.a. 459-480;
D
SNPSALRETVVEVPQVTWEDIG) stabilized the D1 domain. Differential scanning fluorimetry
TE
revealed enhanced stability of the ND1 domain with the linker peptide (i.e., a 13 oC higher melting
temperature). Melting temperatures: ND1 (Tm 41.9 oC, 55.7 oC), ND1L (Tm 55.5 oC) and full-length
AC
CE
P
p97 (Tm 59.7 oC, 66.0 oC).
Figure 2. Analysis of nucleotide binding to p97. (A, B) SPR sensorgrams for ADP (0-16.67 M,
3-fold dilutions) binding to the ND1 (A) and ND1L (B) domains of p97. Response data were fit to a
one-site kinetic model (orange lines) to obtain kinetic binding parameters. (C) SPR sensorgrams
for ADP (0-150 M, 3-fold dilutions), binding to full-length p97. Data were fit to a two-site kinetic
binding model to obtain KD values for the two ATPase domains in p97.
Figure 3. Steady state kinetic analyses of human p97 ATPase activity. (A) ATPase activities of
WT and Walker A and B mutants for D1 (D1-K251A; D1-E305Q) and D2 (D2-K524A; D2-E578Q),
with or without 0.01% Triton X-100 at 200 M ATP. Blue lettering indicates the ATPase active
domain in each protein. Red lettering indicates the Walker A mutant. Green lettering indicates the
Walker B mutant. (B) and (C) Michaelis-Menten plots of ATP hydrolysis for WT and mutant p97.
From (D) to (F), black font indicates the ATPase active domain in each protein. (D) WT p97
hydrolyzes 7.48  0.01 ATP molecules per minute per monomer (turnover number, k cat, min-1). The
21
ACCEPTED MANUSCRIPT
Walker A mutation of D2 (D2-K524A) decreases kcat 22-fold. (E) The apparent Michaelis-Menten
constant, Km, of WT p97 is 287  10 M. The Walker B mutation of D2 (D2-E578Q) decreases Km
50-fold. (F) The catalytic efficiency (kcat/Km) for WT p97 is 0.026 min-1 M-1. A 2-fold decrease for
PT
the Walker A mutation of D1 (D1-K251A), a 3-fold increase for the Walker B mutation of D1 (D1-
RI
E305Q), a 15-fold decrease for the Walker A mutation of D2 (D2-K524A), and a 14-fold increase
SC
for the Walker B mutation of D2 (D2-E578Q) together suggest that D1 is a catalytically competent
ATPase, when D2 is able to bind to nucleotides.
NU
Figure 4. ML240 and ML241 selectively inhibit the D2 domain. (A) Structures of DBeQ, ML240,
ML241, and NMS-873. (B) to (E) IC50 (M) of WT, D1-E305Q, D1-K251A, D2-E578Q, D2-K524A,
MA
and ND1L for (B) DBeQ, (C) ML240, (D) ML241, and (E) NMS-873 with 200 M ATP or 875 M
ATP.
D
Figure 5. Binding of D2 selective inhibitors does not alter ATP binding to the D1 domain. (A)
TE
Fluorescence anisotropy measurements for BODIPY-FL-ATP binding to WT p97 and Walker A
AC
CE
P
mutants. (B) Binding of the BODIPY-FL-ATP probe in the presence of the D2 specific inhibitors,
ML240 and ML241. Data were determined in triplicate, and errors in binding affinities shown are
standard errors of the mean of the fit.
Figure 6. IBMPFD/ALS p97 mutations affect potency of ML240 and ML241. (A) ATPase
activities of WT p97 and the A232E disease mutant (with Walker B mutations) indicated that the
elevated ATPase activity originated from D2. (B) Inhibition of ATPase activity of WT p97 and eight
disease mutants by ML240 and ML241.
Figure 7. A Major p97 cofactor, p47, decreased potency of ML240 and ML241. IC50 values
(M) for p97 inhibitors against p97-p47 mixtures for (A) DBeQ, (B) ML240, (C) ML241, and (D)
NMS-873.
22
B
Figure 1
AC
CE
P
TE
D
MA
NU
A
SC
RI
PT
ACCEPTED MANUSCRIPT
C
D
23
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
NU
A
24
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
NU
SC
B
25
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
NU
C
Figure 2
26
TE
AC
CE
P
A
D
MA
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
27
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
B
28
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
C
29
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
D
30
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
E
31
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
F
Figure 3
32
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
NU
SC
A
33
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
B
34
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
C
35
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
D
36
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
E
Figure 4
37
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
NU
SC
A
38
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
NU
SC
B
Figure 5
39
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
A
40
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
B
Figure 5
41
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
A
42
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
B
43
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
C
44
NU
SC
RI
PT
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
D
Figure 7
45
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
NU
SC
RI
PT
GRAPHICAL ABSTRACT
46
ACCEPTED MANUSCRIPT
AC
CE
P
TE
D
MA
NU
SC
RI
PT
Highlights
Potent p97 inhibitors are developed, but whether they inhibit D1 activity is unknown.
Inhibitors can differentially target D1, D2, and p97-cofactor complexes.
This represents a step toward domain- and complex-selective p97inhibitors.
ML240 and ML241 are D2-selective inhibitors and p47 influences their potency.
47